You are on page 1of 7

Anna. T. Valota, Ian A. Kinloch, Kostya S. Novoselov, Cinzia Casiraghi, Axel Eckmann, Ernie W.

Hill,^ and
Robert A. W. Dryfe,*

ARTICLE

Electrochemical Behavior of Monolayer


and Bilayer Graphene

School of Chemistry, School of Materials, School of Physics & Astronomy, and ^School of Computer Science, University of Manchester, Oxford Road,
Manchester M13 9PL, United Kingdom

he unique charge transport properties of graphene have attracted enormous interest. Among the many
conceived applications of graphene, there
is also much interest in its use as an electrode material, with energy storage/conversion applications in mind. Examples include
recent reports of graphene-based supercapacitors13 and of the use of graphene
as (almost) transparent electrodes in solar
cells.49 Optimization of such applications
requires complete understanding of the
electron transfer properties of graphene.
However, despite a considerable amount
of eort applied in this direction,1015 the
understanding of the electrochemical activity of graphene remains controversial.16 This
work in fact feeds in to the considerable
debate about the relative activity of the
basal plane and edge planes of (bulk) graphitic samples.1721 Much of the lack of
clarity about the electrochemical properties
of graphene stems from the rather ill-dened
sample preparation of almost all of the earlier reports on graphene electrochemistry:
generally, graphene akes (often a mixture
of monolayer, bilayer and other samples)
are simply dispersed on a conducting substrate, with no attempt made to isolate
individual akes, or to isolate the ake
edges from the basal plane. Furthermore,
the majority of studies to date have obtained
graphene via chemical means, usually by
reduction of graphene oxide, leaving doubts
about the purity of the material.2224 There
are a few notable exceptions to this general
trend. Work on the capacitance of graphene
has used a top-gated (eectively, electrochemical) conguration to bias the graphene/electrolyte interface.25,26 More recently, work has appeared which describes
the electrochemical properties of a single,
masked ake in contact with an aqueous
solution of a ferrocene derivative.27 In the
latter article, the authors noted an increase
VALOTA ET AL.

ABSTRACT Results of a study on the electrochemical properties of exfoliated single and

multilayer graphene akes are presented. Graphene akes were deposited on silicon/silicon oxide
wafers to enable fast and accurate characterization by optical microscopy and Raman spectroscopy.
Conductive silver paint and silver wires were used to fabricate contacts; epoxy resin was employed as
a masking coating in order to expose a stable, well-dened area of graphene. Both multilayer and
monolayer graphene microelectrodes showed quasi-reversible behavior during voltammetric
measurements in potassium ferricyanide. However, the standard heterogeneous charge transfer
rate constant, k, was estimated to be higher for monolayer graphene akes.
KEYWORDS: graphene . electrochemistry . Raman spectroscopy . electron transfer

in electron transfer rate for the oxidation of


the ferrocene derivative on the graphene
sample, compared to values quoted on
bulk graphite (highly oriented pyrolytic
graphite, HOPG). The authors were unable
to measure a rate constant for electron
transfer on mechanically exfoliated graphene, the process was suciently fast that
it was essentially reversible. A nite (slower)
rate constant was determined on graphene
prepared by chemical vapor deposition
methods, this rate was still substantially
higher than that measured on basal plane
HOPG. The enhanced electron transfer kinetics seen on graphene were ascribed to
the ripples present on its surface by the
authors of this work. In this report, we have
directly compared the voltammetric responses on a multilayer graphitic surface,
with the response obtained for monolayer
and bilayer graphene. Ferricyanide, which
has been reported to show poor charge
transfer kinetics on basal plane HOPG,18,19
has been chosen as a model system to
highlight dierences in the behavior of the
solids. Accordingly the rst report of charge
transfer kinetics on a monolayer graphene
sample are presented, conrming that electron transfer kinetics are indeed improved
on the graphene surface, relative to bulk
graphitic materials.
VOL. 5

NO. 11

* Address correspondence to
robert.dryfe@manchester.ac.uk.
Received for review July 29, 2011
and accepted October 5, 2011.
Published online October 05, 2011
10.1021/nn202878f
C 2011 American Chemical Society

88098815

2011

8809
www.acsnano.org

ARTICLE

Figure 1. Schematic diagram of the samples employed in


the present work. On the left, top view of the samples; on
the right, cross section along the black dashed line. Samples
are classied as defect-free monolayers, with all the edges
covered by the masking resin (a, b); defective monolayers
with evident holes (c, d); monolayers with exposed edges
(e, f), and multilayers (g, h).

Figure 2. Optical micrographs of the multilayer sample


before (a) and after (b) voltammetric experiments.

RESULTS
The reduction of aqueous phase ferricyanide was
selected as our model redox process. A reduction
process was chosen, to avoid the risk of oxidation of
the graphene samples. Furthermore, a redox couple
with a relatively low standard reduction potential was
chosen to minimize any interference due to the reduction of the solvent background. The ferricyanide couple is interesting because it has been extensively
investigated on carbon surfaces and is reported to
have slow electron transfer kinetics on the basal plane
of well-dened bulk graphitic phases (specically,
highly oriented pyrolytic graphite, HOPG).19 A schematic diagram showing an overview of the analyzed
samples is presented in Figure 1. Samples may be
classied as defect-free monolayers, with all the edges
covered by the masking resin (Figure 1a,b), defective
monolayers with evident holes (Figure 1c,d), monolayers with exposed edges (Figure 1e,f) and multilayers,
which for the purposes of this diagram means two or
more layers (Figure 1g,h). Optical micrographs of a
multilayer sample (>20 graphene sheets) are shown in
Figure 2. Flakes of natural graphite were produced with
a thickness varying from tens to hundreds of nanometers. AFM topography has previously revealed that
steps and folds generated by cleavage of natural
graphite are usually 1020 nm high, with about one
step edge every 800 nm.28 Such defects are evident on
VALOTA ET AL.

Figure 3. Raman spectra performed on two dierent spots


on the multilayer graphene sample. Excitation wavelength,
633 nm; 50 objective; 0.7 mW power (Renishaw
spectrometer).

the multilayer sample in Figure 2a, before exposure to


the solution. Special attention was paid during masking of the samples in order to expose areas with the
minimum number of defects; however, to date it has
not been possible to achieve a perfect, edge-free
region. However, Figure 2b indicates that no obvious
change in the sample was observed following voltammetry. Raman spectra of the sample are presented in
Figure 3. The voltammetric response observed on our
multilayer sample is entirely consistent with previous
literature reports for the ferri/ferrocyanide couple on
VOL. 5

NO. 11

88098815

8810

2011
www.acsnano.org

ARTICLE
Figure 4. Voltammetric response of multilayer graphene at 5 mV/s (a) in 1 M KCl and 1 mM ferricyanide electrolyte; (b) the
experimental data (black lines) is plotted along with the ideal response for reversible electron transfer (red lines). (c) The
voltammetric data for various scan rates (see panel) is shown for the monolayer sample 1, the defect-free graphene; panel d
applies the analysis performed for the multilayer sample in panel b to the 5 mV/s data from Sample 1. In both panels b and d,
currents are normalized to the steady-state current.

Figure 5. Optical micrographs of monolayer graphene samples. Sample 1 is shown before (a) and after (b) masking; in image
b edges are completely masked. Sample 2 is shown to contain holes, in panel c; the exposed part of sample 3 is triangular,
hence edges are exposed to solution (d).

(bulk) graphite and will be considered as our reference (see Figure 4a).18,19 The limiting, backgroundsubtracted current obtained at a scan rate of 5 mV s1
was 7.5 nA. The predicted diusion-limited current for
VALOTA ET AL.

an inlaid microdisc electrode varies between 6.2 and


9.5 nA, according to29
Ilim 4nFDcr
VOL. 5

NO. 11

88098815

(1)

8811

2011
www.acsnano.org

ARTICLE
Figure 6. Optical image (a) and Raman spectra (b) performed on seven dierent spots on the defect-free monolayer graphene
(Sample 1), after the rst series of electrochemical measurements. Excitation wavelength, 633 nm; 100 objective; 0.7 mW
laser power (Witec spectrometer). The crosses correspond to the locations where the mapping was performed; the colors of
the crosses correspond to the colors of the Raman spectra. Map of the G peak integrated area (c) and of the ratio between the
D peak and G peak integrated areas (d).

where n is the number of electrons exchanged in the


redox reaction (1 in the present case), F is the Faraday
constant, D is the diusion coecient of the electroactive species in solution (reported to vary between
5.37  106 cm2 s1 30 and 8.20  106 cm2 s1 31 for
this solute), c is the bulk concentration of the ferricyanide, and r is the radius of the window to the
solution. The standard electron transfer rate constant, k0, for ferricyanide reduction on the multilayer
graphite surface was calculated, from the procedure
described by Mirkin and Bard,32 to be 7  104 cm s1.
Ferricyanide reduction on graphite has been widely
studied in literature, most notably by McCreery and
co-workers.19,33 For HOPG with very low defect
density, standard electron transfer rate constants
lower than 106 cm s1 were found. However, a 1%
defect density is estimated to cause a 103 factor
increase in k0. 18
The voltammetric responses of various monolayer
graphene samples, which had dierent levels of defect
visible, were investigated using the same redox couple
(ferri/ferrocyanide, also shown in Figure 4). Micrographs of these samples are shown in Figure 5. Monolayer sample 1, shown in Figure 5a,b, contained no
VALOTA ET AL.

visible defects and its edges were completely masked,


a conclusion supported by the fact that the ratio between the intensity of D and G peaks was lower than
0.1 in each point of the Raman map34 (see Figure 6d).
Before the masking process (Figure 5a), the surface
of graphene Sample 1 appeared as homogeneous,
with a characteristic color.35 After masking was completed, the presence of bright dots was revealed by
optical microscopy (Figure 5b). The nature of the
observed dots, transparent to Raman spectroscopy
(see Figure 6a,b) is unclear. However, the appearance
of the dots just after the masking process indicates
possible contamination of the graphene surface by
epoxy particles.
Figure 7a shows the background voltammetric
response of Sample 1 (masked, defect-free). Measurement of the non-Faradaic current density (at 0 V)
as a function of sweep rate gives an estimate of the
total interfacial capacitance per unit area, found to
be 21.3 F cm2 (Figure 7b). In aqueous electrolytes,
the total interfacial capacitance is composed of the
quantum capacitance contribution from the graphene in series with the capacitance of the solution
double layer, which can be resolved into a diuse
VOL. 5

NO. 11

88098815

8812

2011
www.acsnano.org

ARTICLE
Figure 8. Current (normalized to electrode radius) vs
potential response for the graphene monolayer samples
(Samples 1 and 2), a bilayer sample and the multilayer.

Figure 7. Current response of defect-free monolayer graphene (Sample 1) in background electrolyte (3 M aqueous
solution of KCl) at various scan rates (a); plot of current
density measured at 0 V versus scan rate (b). The gradient of
the tted line indicates a capacitance of 21.3 F/cm2.

and a compact component.36 In the case of concentrated electrolytes such as the one used in the
present work (3 M), the capacitance of the diuse
ionic layer is usually large (>100 F cm2),36 therefore its contribution to the total capacitance is
negligible. The capacitance arising from the compact
layer is known to have a value of about 1020 F cm2,36
suggesting that the value quoted above for the total
interfacial capacitance is anomalously high. We note that
capacitance of graphene has previously been quoted as
between 8 and 10 F cm2 in 1 mM NaF aqueous
solution, employing a.c. impedance spectroscopy, hence
further investigation is required in order to clarify the
aqueous phase capacitance behavior of monolayer
graphene.25
Ferricyanide voltammetry at Sample 1 is shown in
Figure 4c. As with the multilayer sample, the limiting
current here (8.5  109 A), calculated via eq 1, is in
reasonable agreement with the experimental value
(9.6  109 A), assuming a ferricyanide diusion
coecient of 5.4  106 cm2 s1.30 However, the
currentpotential response is more reversible than
the multilayer graphite case, see Figure 4b,d, indicating
that the graphene surface, despite its apparent lower
level of defects, acts as an ecient catalyst for electron
transfer from ferricyanide. A k0 value of 1.2  103 cm s1
was found for the monolayer sample (Sample 1),
VALOTA ET AL.

almost twice as high as the standard rate of electron


transfer estimated with the (defect containing)
multilayer.
As a matter of control we have also prepared
samples with a number of defects. For instance, monolayer Sample 2 presented in Figure 5c contains several
holes of ca. 10 m diameter, hence some edge sites
must be in contact with the electrolyte; the exposed
part of monolayer Sample 3 is triangular (Figure 5d),
hence edges are also exposed to solution in this case.
Surprisingly, in view of received wisdom about the role
of defects on electron transfer rates for bulk graphitic
samples, the voltammetric responses of the defective
monolayer samples (Samples 2 and 3) were not markedly dierent from that of the defect-free sample
(Sample 1), at least for the case of ferricyanide reduction. Figure 8 presents the comparative current
potential response, where the current is normalized by
sample radius to account for the dierent exposed
windows of the monolayer samples (c.f., eq 1). The
current data obtained from a bilayer graphene sample
is also presented, indicating that this material also
presents electron transfer kinetics which are more
similar to monolayer graphene than to the multilayer
graphitic material.
It should be noted that the electron transfer kinetics
obtained at the monolayer graphene samples degrade
somewhat over a period of time. Figure 8 also shows
the response from the monolayer Sample 2 (few holes,
see Figure 5c) after a two week exposure to ambient
conditions. The normalized current response is almost
identical to that seen for the multilayer sample. A
further feature we have observed is that defects on
the graphene surface appear to act as nucleation sites
for the fracture of samples under the inuence of the
applied potential. This phenomenon is obviously of
interest and is the focus of ongoing work.
CONCLUSIONS
Monolayer and bilayer samples of graphene are
electroactive and present improved electron transfer
VOL. 5

NO. 11

88098815

8813

2011
www.acsnano.org

EXPERIMENTAL METHODS
Fabrication of Electrodes. Samples of monolayer and bilayer
graphene, and multilayer (>20 graphene sheets) graphite, were
prepared by mechanical exfoliation of natural graphite
supplied by NGS Naturgraphit GmbH. The samples were
transferred to a silicon wafer, covered with a 90 nm thick
thermal oxide layer. Conductive silver paint and silver wires
were used to fabricate contacts. The samples were masked
with an epoxy resin to leave a window of the order of 50 m
in diameter, deliberately exposing either the basal plane of
each sample, or the basal plane and some of its edges. The
precise dimensions of each exposed window were determined
by optical microscopy. The samples were characterized by
Raman spectroscopy, either using a Renishaw spectrometer
(50 objective, 0.7 mW power) or a Witec spectrometer
(100 objective, 0.6 mW power) at an excitation wavelength
of 633 nm.
Chemicals and Electrochemical Apparatus. Voltammetric experiments were performed in aqueous solution using a three
electrode configuration under potentiostatic control (Autolab
PGSTAT30, Utrecht, The Netherlands). The masked graphene/
graphite samples were used as the working electrode, an Ag/
AgCl wire (prepared in-house) was used as the reference
electrode, a Pt gauze was employed as the counter electrode.
Water was obtained from an ELGA PureLab-Ultra purifier
(minimum resistance 18.2 M cm). For the electrolyte solutions,
the redox active salt, K3Fe(CN)6, and the supporting electrolyte,
KCl, were purchased from Sigma-Aldrich and Fisher-Scientific,
respectively, and used as received. The pH of the freshly
prepared ferricyanide electrolyte solution was 5.8 (Hanna Instruments pH meter).
Acknowledgment. We thank the U.K. EPSRC (Grants EP/
I005145/1 and EP/G035954/1), the Royal Society, and the University of Manchester EPS strategic equipment fund for nancial
support. We thank J. Martin for assistance with preliminary
experiments.

REFERENCES AND NOTES


1. Stoller, M. D.; Park, S.; Zhu, Y.; An, J.; Ruo, R. S. GrapheneBased Ultracapacitors. Nano Lett. 2008, 8, 34983502.
2. Wang, Y.; Shi, Z.; Huang, Y.; Ma, Y.; Wang, C.; Chen, M.; Chen,
Y. Supercapacitor Devices Based on Graphene Materials.
J. Phys. Chem. C 2009, 113, 1310313107.
3. Zhang, L. L.; Zhou, R.; Zhao, X. S. Graphene-Based Materials
as Supercapacitor Electrodes. J. Mater. Chem. 2010, 20,
59835992.
4. Kumar, A.; Zhou, C. The Race to Replace Tin-Doped Indium
Oxide: Which Material Will Win? ACS Nano 2010, 4, 1114.
5. Choi, H.; Kim, H.; Hwang, S.; Han, Y.; Jeon, M. Graphene
Counter Electrodes for Dye-Sensitized Solar Cells Prepared
by Electrophoretic Deposition. J. Mater. Chem. 2011, 21,
75487551.
6. Wang, Y.; Chen, X.; Zhong, Y.; Zhu, F.; Loh, K. P. Large Area,
Continuous, Few-Layered Graphene as Anodes in Organic
Photovoltaic Devices. Appl. Phys. Lett. 2009, 9, 063302.
7. Wu, J.; Becerril, H. A.; Bao, Z.; Liu, Z.; Chen, Y.; Peumans, P.
Organic Solar Cells with Solution-Processed Graphene
Transparent Electrodes. Appl. Phys. Lett. 2008, 92, 263302.
8. Hecht, D. S.; Hu, L.; Irvin, G. Emerging Transparent Electrodes Based on Thin Films of Carbon Nanotubes, Graphene, and Metallic Nanostructures. Adv. Mater. 2011,
23, 14821513.
9. Hsieh, C. T.; Yang, B. H.; Lin, J. Y. One- and Two-Dimensional
Carbon Nanomaterials as Counter Electrodes for DyeSensitized Solar Cells. Carbon 2011, 49, 30923097.

VALOTA ET AL.

present on the monolayer make little dierence to the


voltammetric response of the samples.

10. Chen, D.; Tang, L.; Li, J. Graphene-Based Materials in


Electrochemistry. Chem. Soc. Rev. 2010, 39, 31573180.
11. Dan, Y.; Lu, Y.; Kybert, N. J.; Luo, Z.; Johnson, A. T. C. Intrinsic
Response of Graphene Vapor Sensors. Nano Lett. 2009, 9,
14721475.
12. Robinson, J. T.; Perkins, F. K.; Snow, E. S.; Wei, Z.; Sheehan,
P. E. Reduced Graphene Oxide Molecular Sensors. Nano
Lett. 2008, 8, 31373140.
13. Alwarappan, S.; Erdem, A.; Liu, C.; Li, C. Z. Probing the
Electrochemical Properties of Graphene Nanosheets for
Biosensing Applications. J. Phys. Chem. C 2009, 113, 8853
8857.
14. Tang, L.; Wang, Y.; Li, Y.; Feng, H.; Lu, J.; Li, J. Preparation,
Structure, and Electrochemical Properties of Reduced
Graphene Sheet Films. Adv. Funct. Mater. 2009, 19, 2782
2789.
15. Zhou, M.; Zhai, Y.; Dong, S. Electrochemical Sensing and
Biosensing Platform Based on Chemically Reduced Graphene Oxide. Anal. Chem. 2009, 81, 56035613.
16. Goh, M. S.; Pumera, M. The Electrochemical Response of
Graphene Sheets Is Independent of the Number of Layers
from a Single Graphene Sheet to Multilayer Stacked
Graphene Platelets. Chem. Asian J. 2010, 5, 23552357.
17. Rice, R. J.; McCreery, R. L. Quantitative Relationship between Electron Transfer Rate and Surface Microstructure
of Laser-Modied Graphite Electrodes. Anal. Chem. 1989,
61, 16371641.
18. Robinson, R. S.; Sternitzke, K.; McDermott, M. T.; McCreery,
R. L. Morphology and Electrochemical Eects of Defects on
Highly Oriented Pyrolytic Graphite. J. Electrochem. Soc.
1991, 138, 24122418.
19. Cline, K. K.; McDermott, M. T.; McCreery, R. L. Anomalously
Slow Electron Transfer at Ordered Graphite Electrodes:
Inuence of Electronic Factors and Reactive Sites. J. Phys.
Chem. 1994, 98, 53145319.
20. Banks, C. E.; Compton, R. G. Edge Plane Pyrolytic Graphite
Electrodes in Electroanalysis: An Overview. Anal. Sci. 2005,
21, 12631268.
21. Edwards, M. A.; Bertoncello, P.; Unwin, P. R. Slow Diusion
Reveals the Intrinsic Electrochemical Activity of Basal
Plane Highly Oriented Pyrolytic Graphite Electrodes.
J. Phys. Chem. C 2009, 113, 92189223.
22. Shan, C.; Yang, H.; Song, J.; Han, D.; Ivaska, A.; Niu, L. Direct
Electrochemistry of Glucose Oxidase and Biosensing for
Glucose Based on Graphene. Anal. Chem. 2009, 81, 2378
2382.
23. Wang, J.; Yang, S.; Guo, D.; Yu, P.; Li, D.; Ye, J.; Mao, L.
Comparative Studies on Electrochemical Activity of Graphene Nanosheets and Carbon Nanotubes. Electrochem.
Commun. 2009, 11, 18921895.
24. Wu, J. F.; Xu, M. Q.; Zhao, G. C. Graphene-Based Modied
Electrode for the Direct Electron Transfer of Cytochrome C
and Biosensing. Electrochem. Commun. 2010, 12, 175177.
25. Xia, J.; Chen, F.; Li, J.; Tao, N. Measurement of the Quantum
Capacitance of Graphene. Nat. Nanotechnol. 2009, 4,
505509.
26. Das, A.; Pisana, S.; Chakraborty, B.; Piscanec, S.; Saha, S. K.;
Waghmare, U. V.; Novoselov, K. S.; Krishnamurthy, H. R.;
Geim, A. K.; Ferrari, A. C. Monitoring Dopants by Raman
Scattering in an Electrochemically Top-Gated Graphene
Transistor. Nat. Nanotechnol. 2008, 3, 210215.
27. Li, W.; Tan, C.; Lowe, M. A.; Abru~
na, H.; Ralph, D. C. Electrochemistry of Individual Monolayer Graphene Sheets. ACS
Nano 2011, 5, 22642270.
28. Gu, J.; Leng, Y. Study of the Surface Topography of
Graphite Materials Using Atomic Force Microscopy. Carbon 1999, 37, 991994.

VOL. 5

NO. 11

88098815

ARTICLE

kinetics, for the case of ferricyanide reduction, compared to the basal plane graphite substrates. Defects

8814

2011
www.acsnano.org

ARTICLE

29. Bard, A. J.; Faulkner, L. R. Electrochemical Methods: Fundamentals and Applications; Wiley: New York, 1980.
30. Plana, D.; Jones, F. G. E.; Dryfe, R. A. W. The Voltammetric
Response of Bipolar Cells: Reversible Electron Transfer.
J. Electroanal. Chem. 2010, 646, 107113.
31. Winkler, K. The Kinetics of Electron Transfer in Redox
System on Platinum Standard-Size and Ultramicroelectrodes. J. Electroanal. Chem. 1995, 388, 151159.
32. Mirkin, M. V.; Bard, A. J. Simple Analysis of Quasi-Reversible
Steady-State Voltammograms. Anal. Chem. 1992, 64,
22932302.
33. Kneten, K. R.; McCreery, R. L. Eects of Redox System
Structure on Electron-Transfer Kinetics at Ordered Graphite and Glassy Carbon Electrodes. Anal. Chem. 1992, 64,
25182524.
34. Casiraghi, C.; Hartschuh, A.; Qian, H.; Pliscanec, S.; Georgia,
C.; Fasoli, A.; Novoselov, K. S.; Basko, D. M.; Ferrari, A. C.
Raman Spectroscopy of Graphene Edges. Nano Lett. 2009,
9, 14331441.
35. Blake, P.; Hill, E. W.; Castro Neto, A. H.; Novoselov, K. S.;
Jiang, D.; Yang, R.; Booth, T. J.; Geim, A. K. Making Graphene
Visible. Appl. Phys. Lett. 2007, 91, 063124.
36. Randin, J.-P.; Yeager, E. Dierential Capacitance Study of
Stress-Annealed Pyrolytic Graphite Electrodes. J. Electrochem. Soc. 1971, 118, 711714.

VALOTA ET AL.

VOL. 5

NO. 11

88098815

8815

2011
www.acsnano.org

You might also like